In this study, we present a one-dimensional tight-binding model designed to explore the impact of electric fields on an incommensurate quantum system. We specifically focus on the Aubry–André–Harper model, a quasiperiodic model known to exhibit a metal–insulator transition at a critical potential value of λc = 2. This model combines Anderson and Aubry–André–Harper localization phenomena in a quantum system, leading to intriguing effects on the lattice band structure upon the application of an electric field F to the Aubry–André–Harper potential. Our investigation reveals that by choosing a specific value for the applied electric field, it becomes feasible to generate effective massless Dirac fermions within our Aubry–André–Harper system. Furthermore, we note that the extension or localization of the massless particle wave function is contingent upon the potential strength value λ within our incommensurate model. Importantly, our findings highlight the potential for detecting this intriguing phenomenon through experimental means.

Since Anderson’s groundbreaking work,1 the investigation into quantum localization has remained a cornerstone of condensed matter physics. Despite extensive research into localization phenomena over decades, numerous intriguing questions persist in this domain. The Anderson model and the Aubry–André–Harper (AAH) incommensurate potential model continue to captivate physicists’ attention. While the Anderson model introduces disorder through random amplitudes in lattice on-site positions, the AAH model2–4 stands out by imposing disorder derived from a potential function lacking commensurability with the lattice.

Localization effects in the AAH model stem from the intricate interplay between destructive interference during multiple scattering processes of a particle traversing an incommensurate medium and the spectral characteristics of its Schrödinger equation.5 Recent years have seen a surge in interest in this area, primarily fueled by innovative experimental studies using optical lattices and cold atomic gases.6–9 

In particular, the AAH model in optical lattices has posed an intriguing challenge, shedding light on the elusive metal–insulator transition that occurs with variations in the periodic potential’s intensity.10–13 Researchers have explored delocalization phenomena within the AAH potential, notably utilizing ultracold bosons.14,15 Recent studies have contributed to understanding the interaction range in both Anderson and AAH systems.11 Xu et al.13 revealed a Griffiths-like regime in interacting AAH models. Investigations into the nonlinear response of interacting bosons in quasiperiodic potentials have also provided valuable insights.6 In addition, the concept of many-body localization within incommensurate systems has been examined by Huang et al.9 A recent report on hopping tunneling through a slowly varying potential is also noteworthy.16 

It is important to note that experimental demonstration of an AAH transition in real materials remains a significant challenge. While the detection of an AAH transition has been accomplished through optical measurement techniques using ultracold atoms,2 solid-state superlattices have emerged as a promising platform for demonstrating this transition in real crystals. Solid-state superlattices involve the alternating growth of two different materials, with one acting as a potential barrier and the other as a quantum well.17,18 These structures have garnered considerable attention due to their potential applications in novel devices.19–22 Recent studies have suggested the existence of phonon Anderson localization in Si/Ge aperiodic superlattices, evident through their thermal conductivity behaviors.19–21 

Moreover, the captivating Klein paradox, an extraordinary phenomenon emerging from relativistic physics and featuring massless Dirac fermions, has been a focal point in recent scientific discourse. In the groundbreaking theoretical investigation by Katsnelson et al.,23 the researchers showcased an impeccable transmission for massless Dirac fermions approaching a one-dimensional (1D) potential barrier within a solid-state apparatus. This discovery stands as a solid-state analog24–30 to the manifestation of the Klein paradox, an intriguing relativistic tunneling effect initially postulated by Klein in the context of 1D potentials.31 The implication is that a complete confinement of massless Dirac fermions within a solid-state crystal through a 1D potential barrier becomes unattainable.

Theoretical propositions have surfaced regarding the containment of massless Dirac fermions in graphene, suggesting the utilization of circular graphene pn junctions.32–36 Experimental endeavors, employing scanning tunneling microscopy (STM), have successfully revealed that massless Dirac fermions can undergo a temporary entrapment within circular pn junctions. The wave functions of these charge carriers are confined, adopting quantified modes through continuous reflection near the interface of closed-circle potential barriers.37–40 These outcomes not only confirm the anomalous anisotropic transmission of massless Dirac fermions at pn junction boundaries, known as Klein tunneling in graphene,23 but also provide insights at the atomic scale.

With these considerations in mind, our objective is to evaluate the applied bias effects on an incommensurate potential (Fig. 1). To date, there have been no measurements of carrier transport using incommensurate potentials in real materials. A pivotal aspect of this research area revolves around comprehending electric field effects on quasiperiodic systems. In our tight-binding model, we combine Anderson-type localization, produced by an external bias, and AAH-type localization, given by the incommensurate potential. Intriguingly, we also uncover a novel phenomenon: the presence of massless Dirac fermions in an AAH system at certain values of an applied electric field.

FIG. 1.

Aubry–André–Harper (AAH) and periodic potentials plotted against the lattice site n in the absence of an external electric field. The parameters utilized are ϕ = 0 and λ = 2.0 for the potential described in Eq. (2). For the incommensurate potential, α = 0.618 03 (solid line) is employed, while in the periodic case, α = 0.600 00 (dotted line) is considered. The arrow indicates the incommensurate potential peaks, displaying a non-periodic behavior. In the case of the periodic potential (dotted line), the potential peaks align neatly along a straight line, as highlighted by the dashed line.

FIG. 1.

Aubry–André–Harper (AAH) and periodic potentials plotted against the lattice site n in the absence of an external electric field. The parameters utilized are ϕ = 0 and λ = 2.0 for the potential described in Eq. (2). For the incommensurate potential, α = 0.618 03 (solid line) is employed, while in the periodic case, α = 0.600 00 (dotted line) is considered. The arrow indicates the incommensurate potential peaks, displaying a non-periodic behavior. In the case of the periodic potential (dotted line), the potential peaks align neatly along a straight line, as highlighted by the dashed line.

Close modal

In addition, we have found that fermion eigenvectors reflect the specific characteristics of the quasiperiodic eigenstates. They can be extended or localized in the AAH potential. We will show the possibility of detecting such massless carriers by means of Klein tunneling through a solid-state device. The existence of Dirac carriers in an incommensurate potential can also be employed to demonstrate the AAH metal–insulator transition in a solid-state material, which has not yet been detected. Our overarching goal is to underscore the critical role of field effects in advancing our understanding of experiments involving incommensurate quantum systems. Through these insights, we aspire to make a substantial contribution to the fundamental understanding of the mechanisms governing the massless electron behavior in solid-state superlattices and real crystals.

The systems under investigation are characterized by a tight-binding equation, a widely used model in this research field. This equation describes the movement of an individual particle along a one-dimensional lattice, where each lattice site connects to its immediate neighbors through a matrix element denoted as t. The tight-binding equation is expressed as follows:
(1)
where un represents the amplitude of the carrier wave function at the nth lattice site, Vn denotes the on-site potential, and F represents the applied electric field, calculated as F = V/aN, in which a stands for the lattice parameter, V is the bias applied by the two lattice end sites, and N is the number of lattice sites. For simplicity, we have set the value of the nearest-neighbor hopping parameter t to 1. Furthermore, the fundamental electric charge e is considered as 1 for convenience.

It is vital to emphasize that the particular form of the on-site potential Vn in Eq. (1) entirely defines the physical model under examination. For instance, a periodic Vn corresponds to the well-known Bloch case, while a random Vn corresponds to the Anderson localization model.1 In periodic potentials, all states extend indefinitely, whereas random potentials Vn lead to entirely localized states.10,11 The incommensurate scenario lies between periodic and random cases. In an AAH system, Vn is periodic, but the period is not a rational multiple of the underlying lattice.12–15 

This study delves into investigating incommensurate potentials using a tight-binding model. In the original AAH model, the potential is mathematically defined as
(2)
where α is an irrational number (α=1+5), ϕ represents a phase, and λ denotes the potential strength. Notably, Vn exhibits a metal–insulator transition when λ equals 2.

Consistent with prior work,12 we have selected α=2/(51)=0.61803, characterizing quasiperiodicity in certain quasicrystals in the incommensurate case12 (see Figs. 1 and 2). When α is a real number, strictly speaking, the potential is (almost) periodic given that n is an integer and N is a finite number. In this work, we refer to the potential as periodic when α = 0.6, and as quasiperiodic, or Aubry–André–Harper (AAH), or incommensurate when α = 0.618 032. Figures 1 and 2 illustrate both periodic and quasiperiodic potentials. Our findings are corroborated by averaging across various ϕ values.

FIG. 2.

Aubry–André–Harper (AAH) and periodic potentials plotted against the lattice site n under certain applied bias conditions. An external bias of V = 6 has been applied between the lattice end sites. The parameters utilized are ϕ = 0 and λ = 2.0 for the potential described in Eq. (2). For the incommensurate potential, α = 0.618 03 (solid line) is employed. In the case of the periodic potential, α = 0.600 00 (dotted line) is considered. The arrow indicates the incommensurate potential peaks, displaying an aperiodic behavior. For the periodic potential (dotted line), the potential peaks align neatly along a straight line, as highlighted by the dashed line.

FIG. 2.

Aubry–André–Harper (AAH) and periodic potentials plotted against the lattice site n under certain applied bias conditions. An external bias of V = 6 has been applied between the lattice end sites. The parameters utilized are ϕ = 0 and λ = 2.0 for the potential described in Eq. (2). For the incommensurate potential, α = 0.618 03 (solid line) is employed. In the case of the periodic potential, α = 0.600 00 (dotted line) is considered. The arrow indicates the incommensurate potential peaks, displaying an aperiodic behavior. For the periodic potential (dotted line), the potential peaks align neatly along a straight line, as highlighted by the dashed line.

Close modal
We introduce the normalized density of states, denoted as ρ(E),
(3)
where N represents the number of sites, and the density is normalized to 1. To calculate the density of states, we approximate δ(EEj) using a normalized Gaussian function,41–43 
(4)
In our study, we set σ to σ = 0.15. The Ej values represent the obtained eigenvalues, with j belonging to the set {1, 2, …, N}.
The Lyapunov exponent γ(E) is a widely used parameter for characterizing extended and localized eigenstates in solid state physics. It is defined as the inverse of the localization length of the quantum wave function,41–43 providing insights into the asymptotic behavior of system states,
(5)
where N denotes the number of lattice sites. For extended states, γ(E) = 0, whereas for a localized eigenstate with energy E, γ(E) > 0.
For each eigenvector |un > = {a1, …, an}, we define its averaged position in the lattice as
(6)
Its dispersion σn can be calculated as
(7)
Following a similar approach to Ref. 44, we have calculated the dispersion of each state. We assume that each eigenvector takes the following form:
(8)
Here, we set the position of the packet’s center at ⟨n⟩. For a wave packet prepared in this manner, we compute the dispersion using the following formula:
(9)
This is normalized to ensure that σp = 1 when the eigenstates are uniformly extended (i.e., when un=1/N for all sites n).44 Thus, we have
(10)
This equation provides a measure of the average extension of the eigenstates, where σp = 1 signifies that most eigenstates extend throughout the lattice, whereas σp = 0 indicates a strong localization for all eigenstates.
In the absence of an external electric field, we solved Eq. (1) considering the potential outlined by Eq. (2). Figure 3 illustrates the obtained eigenenergies vs eigenvalue number n, ordered by their energy. We investigated two different potential values, λ = 0.5 and λ = 2.0. For low potential values (λ = 0.5), bandgaps are observed in Fig. 3. The energies align with a typical dispersion relation,
(11)
This representation is typical for a tight-binding model, with k being the crystal momentum, ϑ being a phase, E0 being a constant, and A being the energy band amplitude. Figure 3 indicates that as λ reaches its critical value (λ = 2), the eigenvalues tend to cluster at certain energies.
FIG. 3.

Plot of eigenenergies as a function of eigenvalue number n for two distinct potentials. For both cases, ϕ = 0 and the potentials described in Eq. (2) are employed with λ values set to λ = 0.5 and λ = 2.0. An applied bias of V = 0 is considered. The shaded areas indicate the bandgaps present in the λ = 0.5 Aubry–André–Harper (AAH) potential.

FIG. 3.

Plot of eigenenergies as a function of eigenvalue number n for two distinct potentials. For both cases, ϕ = 0 and the potentials described in Eq. (2) are employed with λ values set to λ = 0.5 and λ = 2.0. An applied bias of V = 0 is considered. The shaded areas indicate the bandgaps present in the λ = 0.5 Aubry–André–Harper (AAH) potential.

Close modal

Figures 4 and 5 portray the density of states and the Lyapunov exponent as a function of energy for an applied bias of V = 0. The parameters λ are set to λ = 0.5 and λ = 2.0, and the energy reference point is placed at the band center. Across various energy strengths, the potential structure described by Eq. (2) yields three energy bands within the lattice. The eigenvalues cluster in the upper, central, and lower regions, generating the peaks depicted in Figs. 4 and 5 within the density of states. In these eigenvalue clusters, the Lyapunov exponent reaches its minimum values, indicating extended states. Remarkably, no significant disparities are evident between the density of states and Lyapunov exponent plots.

FIG. 4.

Plot of the density of states (solid line) against energy for an Aubry–André–Harper (AAH) potential, alongside the Lyapunov exponent graph (dotted-dashed line). The parameters used include ϕ = 0 and λ = 0.5, referring to the potential in Eq. (2). The applied bias is set to 0. In this instance, σ = 0.015, as per Eq. (4). The shaded regions highlight the bandgaps present in the AAH potential with a strength of λ = 0.5.

FIG. 4.

Plot of the density of states (solid line) against energy for an Aubry–André–Harper (AAH) potential, alongside the Lyapunov exponent graph (dotted-dashed line). The parameters used include ϕ = 0 and λ = 0.5, referring to the potential in Eq. (2). The applied bias is set to 0. In this instance, σ = 0.015, as per Eq. (4). The shaded regions highlight the bandgaps present in the AAH potential with a strength of λ = 0.5.

Close modal
FIG. 5.

Plot of the density of states (solid line) against energy for an Aubry–André–Harper (AAH) potential, along with the Lyapunov exponent graph (dotted-dashed line). The parameters used are ϕ = 0 and λ = 2.0, referring to the potential described in Eq. (2). The applied bias is set to 0. The shaded regions highlight the bandgaps present in the AAH potential at λ = 2.0.

FIG. 5.

Plot of the density of states (solid line) against energy for an Aubry–André–Harper (AAH) potential, along with the Lyapunov exponent graph (dotted-dashed line). The parameters used are ϕ = 0 and λ = 2.0, referring to the potential described in Eq. (2). The applied bias is set to 0. The shaded regions highlight the bandgaps present in the AAH potential at λ = 2.0.

Close modal

Figure 6 illustrates the dispersion average σp as a function of potential strength λ, following Eq. (10). We considered different applied biases, V = 0, V = 0.1, and V = 0.2. At V = 0, as λ reaches its critical point (λ = 2), an AAH metal–insulator transition occurs within the incommensurate model. For λ > 2, σp tends toward zero, indicating highly localized eigenstates. Conversely, at λ < 2, σp remains nearly equal to 1, signifying predominantly extended eigenvectors. Figure 6 emphasizes the feasibility of detecting the AAH transition using the proposed σp indicator.

FIG. 6.

Graph illustrating the averaged dispersion σp plotted as a function of potential strength λ for various applied biases. In the case of V = 0, a distinct metal–insulator transition is observable at λ = 2.0.

FIG. 6.

Graph illustrating the averaged dispersion σp plotted as a function of potential strength λ for various applied biases. In the case of V = 0, a distinct metal–insulator transition is observable at λ = 2.0.

Close modal

We would like to emphasize that the AAH potential exhibits the characteristic of being self-dual, with the Hamiltonian maintaining the same form in both momentum and space representations. Consequently, our tight-binding model possesses a self-duality point at λ = 2, delineating the boundary between the localized phase, where all states are confined, and the delocalized phase, where all states extend. At this transition point, the energy spectrum takes on a fractal nature. Nevertheless, the introduction of an electric field disrupts the system’s self-duality. We posit that the fractal behavior of the critical states at the transition point diminishes as F is increased.

Furthermore, the applied bias induces Wannier–Stark localization in the eigenfunctions. Consequently, the initially extended eigenfunctions from the delocalized phase transition into localized states, with the states in the already localized phase exhibiting an even greater confinement. This effect is vividly illustrated in Fig. 6, which demonstrates a decrease in σp as V is increased. For cases where λ > 2, the eigenstates show a strong localization when V > 0. In addition, at λ < 2, observable Wannier–Stark localization effects are evident in the eigenvectors.

In Fig. 6, σp exhibits a sharp behavior in the absence of an external electric field. Conversely, the averaged dispersion becomes smoother when V > 0. We attribute the sharp behavior to the presence of a self-duality point in the model at λ = 2 and V = 0. The self-duality of the model is disrupted by the introduction of the electric field.

Figure 7 depicts the average position, ⟨n⟩, within the lattice as a function of position n for λ = 0.5 and λ = 2.5. At λ = 0.5, the ⟨n⟩ values are concentrated around the midpoint of the lattice, approximately ⟨n⟩ ∼ 250 for a lattice of size N = 500. This outcome aligns with the prevalence of extended states at λ = 0.5, where the probability of finding a carrier is uniform across all sites, resulting in an average position of approximately ⟨n⟩ ∼ 250, the central site of the tight-binding chain.

FIG. 7.

Graph illustrating the averaged position ⟨n⟩ plotted against the lattice site n for both λ = 0.5 and λ = 2.5. The parameters employed here are ϕ = 0 and F = 0 for the potential specified in Eq. (2). In the case of λ = 0.5, the average position ⟨n⟩ remains ∼250 in the majority of instances.

FIG. 7.

Graph illustrating the averaged position ⟨n⟩ plotted against the lattice site n for both λ = 0.5 and λ = 2.5. The parameters employed here are ϕ = 0 and F = 0 for the potential specified in Eq. (2). In the case of λ = 0.5, the average position ⟨n⟩ remains ∼250 in the majority of instances.

Close modal

Conversely, at λ = 2.5, the ⟨n⟩ values are dispersed across all lattice sites, reflecting a strong localization of most eigenstates. Unlike the λ = 0.5 case, there is no privileged value for ⟨n⟩ at λ = 2. For each distinct eigenvector of the system at λ = 2.5, a different ⟨n⟩ value is obtained. This provides a clear visualization of the AAH transition through the ⟨n⟩ indicator.

Now, considering the effects of an electric field (V > 0) on the system, Fig. 6 illustrates the dispersion average σp for small applied biases. The metal–insulator transition at λc shifts to a lower λ value if V > 0. At V = 0, the strongly localized regime (σp ∼ 0) is reached at λ = 2.5. However, with V = 0.2, Fig. 6 shows that σp ∼ 0 at λ = 2.1. Both Anderson and AAH localization models are present in our quantum system when F > 0, resulting in an increased localization of carrier wave functions within the incommensurate lattice.

For deeper insights into this phenomenon, Fig. 8 displays the density of states and the Lyapunov exponent vs energy at λ = 2.0 and V = 1.55. Three peaks in the density of states correspond to the three energy bands observed in the V = 0 case. However, the energy gaps of the AAH potential vanish at V = 1.55. The arrows in Fig. 8 indicate the positions of the energy bandgaps observed when the applied bias is null. Notably, the Lyapunov exponent does not vanish along the energy axis, indicating a localization of all system eigenvectors at λ = 2.0 and V = 1.55.

FIG. 8.

Density of states (solid line) plotted against energy for an AAH potential, accompanied by the Lyapunov exponent graph (dotted-dashed line). The parameters employed include ϕ = 0 and λ = 2.0 for the potential as described in Eq. (2). The applied bias has been taken to be equal to 1.55. The arrows indicate the position of the energy band crossings.

FIG. 8.

Density of states (solid line) plotted against energy for an AAH potential, accompanied by the Lyapunov exponent graph (dotted-dashed line). The parameters employed include ϕ = 0 and λ = 2.0 for the potential as described in Eq. (2). The applied bias has been taken to be equal to 1.55. The arrows indicate the position of the energy band crossings.

Close modal

At λ = 2.0, Fig. 9 displays the band structure of our AAH lattice for two different electric fields, V = 0 and V = 1.55. Band crossings in the energy bands occur at the critical value Vc = 1.55. For applied biases lower than Vc, Dirac cones in the energy bands are not observed. The arrows in Fig. 9, plotted also in Fig. 8, indicate the positions of the band crossings. At V = 0, the eigenvalues are grouped into clusters, as depicted in Fig. 9.

FIG. 9.

Band structure of a 1D tight-binding model at λ = 2.0. We have considered two different applied biases, V = 0 and V = 1.55. The arrows indicate the position of the Dirac crossings.

FIG. 9.

Band structure of a 1D tight-binding model at λ = 2.0. We have considered two different applied biases, V = 0 and V = 1.55. The arrows indicate the position of the Dirac crossings.

Close modal

To illuminate this phenomenon further, Fig. 10 illustrates the formation of the upper Dirac cone of Fig. 9. Three different electric fields, V = 0.00, V = 1.00, and V = 1.55, are considered. The arrow indicates the band crossing, and the two straight lines denote the position of the Dirac cone. The central and upper energy bands do not converge if V < Vc.

FIG. 10.

Eigenenergies vs crystal momentum at three different applied fields: V = 0.0, V = 1.0, and V = 1.55. A formation of a Dirac cone is indicated by two straight lines and an arrow. The solid circles represent V = 1.55, the open circles denote V = 1.0, and the crosses represent V = 0.0.

FIG. 10.

Eigenenergies vs crystal momentum at three different applied fields: V = 0.0, V = 1.0, and V = 1.55. A formation of a Dirac cone is indicated by two straight lines and an arrow. The solid circles represent V = 1.55, the open circles denote V = 1.0, and the crosses represent V = 0.0.

Close modal

To explore if this effect is specific to the chosen potential value (λ = 2.0), Figs. 11 and 12 demonstrate the results for a higher potential (λ = 2.5). In addition, Figs. 13 and 14 exhibit the calculations for a lower λ value (λ = 1.0). In Figs. 1114, the arrows mark the positions of the Dirac crossings. Notably, extended states are observed at the band center in Fig. 13, where the Lyapunov exponent reaches 0 at E = 0. The calculated critical values for the applied bias Vc vary with changes in λ. At λ = 2.5, Vc equals Vc = 1.65, and Vc = 0.95 for the λ = 1.0 case. Figure 15 depicts the calculated critical value for the applied bias Vc vs λ. As the potential strength λ increases in our AAH model, the obtained Vc values also increase.

FIG. 11.

Density of states (solid line) plotted against energy for an AAH potential, along with the Lyapunov exponent graph (dotted-dashed line). The parameters utilized are ϕ = 0 and λ = 2.5 for the potential described in Eq. (2). An applied bias of 1.3 is considered. The arrows highlight the positions of the Dirac cones.

FIG. 11.

Density of states (solid line) plotted against energy for an AAH potential, along with the Lyapunov exponent graph (dotted-dashed line). The parameters utilized are ϕ = 0 and λ = 2.5 for the potential described in Eq. (2). An applied bias of 1.3 is considered. The arrows highlight the positions of the Dirac cones.

Close modal
FIG. 12.

Band structure of a 1D tight-binding model at λ = 2.5. We have considered two different applied biases, V = 0 and V = 1.65. The arrows indicate the position of the Dirac crossings.

FIG. 12.

Band structure of a 1D tight-binding model at λ = 2.5. We have considered two different applied biases, V = 0 and V = 1.65. The arrows indicate the position of the Dirac crossings.

Close modal
FIG. 13.

Density of states (solid line) plotted against energy for an AAH potential, accompanied by the Lyapunov exponent graph (dotted-dashed line). The parameters employed include ϕ = 0 and λ = 1.0 for the potential as described in Eq. (2). The applied bias has been taken to be equal to 0.95. The arrows indicate the position of the band crossings.

FIG. 13.

Density of states (solid line) plotted against energy for an AAH potential, accompanied by the Lyapunov exponent graph (dotted-dashed line). The parameters employed include ϕ = 0 and λ = 1.0 for the potential as described in Eq. (2). The applied bias has been taken to be equal to 0.95. The arrows indicate the position of the band crossings.

Close modal
FIG. 14.

Band structure of a 1D tight-binding model at λ = 1.0. We have considered two different applied biases, V = 0 and V = 0.95. The arrows indicate the position of the Dirac cones.

FIG. 14.

Band structure of a 1D tight-binding model at λ = 1.0. We have considered two different applied biases, V = 0 and V = 0.95. The arrows indicate the position of the Dirac cones.

Close modal
FIG. 15.

Critical applied bias Vc vs potential strength λ.

FIG. 15.

Critical applied bias Vc vs potential strength λ.

Close modal

To verify the robustness of the results under variations in key parameters of the model, such as V and α, we calculated Vc for different α′ values, N = 500. Specifically, at α′ = α + 0.1 and λ = 2, we obtained Vc = 1.58. In addition, for α′ = α − 0.1 and λ = 2, Vc = 1.52 was found.

The arrows depicted in Figs. 814 pinpoint the band crossings in the quantum system, situated at the center of the Dirac cones, for various critical applied bias values. The fermions’ energy dispersion is dictated by these Dirac cones, governed by a Dirac Hamiltonian,
(12)
where σ̂ represents the pseudospin matrix with components given by Pauli matrices, p̂ denotes the momentum operator, and vF stands for the effective spin of the system, representing the Fermi velocity.

The Hamiltonian acts on states represented by the two-component spinor Ψ=[ψa,ψb]T, where ψa and ψb correspond to the envelope functions associated with the probability amplitudes. The low-dimensional spectrum of free carriers is E = ±vFk, with k denoting the wave vectors around the cones at the Brillouin zones. The signs + and − refer to the electron and hole bands, respectively. It is noteworthy that the wave function of these massless particles can be either extended or localized based on the λ value. The generated fermions mirror the localization characteristics of the incommensurate potential eigenfunctions.

To elucidate this aspect, Fig. 16 displays the plots of two eigenvectors for both cases. These eigenvectors are accompanied by Gaussian functions computed using Eq. (8). It becomes evident that at λ = 1.5, the state in the incommensurate model is localized, represented by slender Gaussian functions with minimal σ dispersion values. Conversely, the AAH model demonstrates an extended state at λ = 0.5, illustrated by broader Gaussian curves.

FIG. 16.

Plot of eigenvectors as a function of lattice site n for our AAH model. Each eigenvector corresponds to a different massless Dirac particle. We have considered two strength potentials, λ = 1.5 and λ = 0.5, respectively. Both eigenvectors have the same eigenvalue number n = 346 but correspond to different potentials. The eigenfunctions are fitted using a Gaussian function described in Eq. (8).

FIG. 16.

Plot of eigenvectors as a function of lattice site n for our AAH model. Each eigenvector corresponds to a different massless Dirac particle. We have considered two strength potentials, λ = 1.5 and λ = 0.5, respectively. Both eigenvectors have the same eigenvalue number n = 346 but correspond to different potentials. The eigenfunctions are fitted using a Gaussian function described in Eq. (8).

Close modal

It is important to note that, through the adjustment of the applied bias, we have the capability to intentionally manipulate the band structure, achieving a true crossing of two energy bands known as a Dirac point. As illustrated in Fig. 10, the Dirac cone, marked by an arrow, is achieved through the crossing of the second and third bands of the system at λ = 2. In the proximity of such crossings, the dynamics can be effectively described by a one-dimensional Dirac equation.45 

In a Dirac cone, the states of electrons and holes are interconnected, exhibiting features similar to the charge-conjugation symmetry observed in quantum electrodynamics. This symmetry arises from the inherent potential symmetry in the system, as particles need to be described by two-component wave functions. These wave functions are crucial for determining the relative contributions of up and down sublattices in the composition of effective particles. The two-component description closely parallels the use of spinor wave functions in quantum electrodynamics. In our condensed-matter system, the carriers effectively emulate Dirac fermions as observed in quantum electrodynamics.23 

In Fig. 10, it is evident that when the electric field is below its critical value (V = 1.0), an energy gap exists between both up- and down-cones. In this scenario, the system exhibits massive fermions rather than massless fermions in the potential. To generate massless fermions in the system, a necessary condition is that the applied bias value reaches Vc = 1.55. It is worth noting that there are materials where external fields are not required to have massless fermions in the quantum system. For instance, graphene possesses energy bands with Dirac cones inherent to its atomic structure, and these cones are naturally generated.

Figure 17 exhibits the positions of the Dirac cones for upper and lower massless fermions concerning the AAH potential strength. As λ increases, the energy gap between the fermions also increases. Furthermore, Fig. 18 portrays the Fermi velocity vF of the massless carriers against λ. Different Vc values have been utilized to compute vF for each potential strength value. The observation indicates that vF escalates with increasing λ.

FIG. 17.

Energy of Dirac crossings vs potential strength λ. Triangles: upper energy band. Squares: lower energy band.

FIG. 17.

Energy of Dirac crossings vs potential strength λ. Triangles: upper energy band. Squares: lower energy band.

Close modal
FIG. 18.

Fermi velocity vs AAH potential strength λ.

FIG. 18.

Fermi velocity vs AAH potential strength λ.

Close modal

In conclusion, we posit that the existence of massless particles in an AAH potential can be experimentally detected. A potential experimental setup involves measuring the tunneling current46–48 through an incommensurate potential when an AAH device is inserted between two distinct electrodes. In principle, if the incommensurate potential is in its insulating phase, where all states are localized, carrier transport through the potential will be impractical. Consequently, the expectation of a finite value for the tunneling current through the device under such conditions may not be feasible.

However, considering the relativistic nature of the band structure within our quantum system, the presence of Dirac fermions allows for their transmission through high potential barriers via the Klein tunneling effect. Therefore, even if the system is in its insulating phase, an anticipated current resulting from Klein tunneling may be expected, contingent upon the generation of massless particles within the system.

Our approach involves utilizing a critical value for the external applied bias Vc to facilitate the creation of massless Dirac fermions at the energy crossings within the band structure. Consequently, even though the system is in an insulating phase, a Klein tunneling current can be generated due to the existence of these massless particles. Such an experimental outcome not only holds the potential to serve as evidence of massless carriers but could also represent a groundbreaking demonstration of an AAH metal–insulator transition in a solid-state device, given that this particular transition has yet to be realized experimentally.

In recent years, various platforms have been employed to investigate the AAH potential. The AAH model has been implemented in photonic crystals, cold-atom systems, and superconducting circuits.49 Cold atoms, being ionizable and susceptible to electric fields, are particularly noteworthy in this context.

The emergence of Dirac cones in an AAH system holds promise for potential applications in new experiments and devices. For instance, when the system is in the localized phase and V < Vc, no tunneling current through the device is expected. However, once the V value reaches Vc, Klein tunneling through the AAH potential becomes feasible, facilitated by the existence of Dirac cones in the energy bands. This opens the possibility of implementing a new resonant tunneling-type device.

From both an experimental and theoretical standpoints, exploring many-body effects in the AAH model would be intriguing. Recent investigations using optical lattices and cold atomic gases have examined many-body localization effects in an AAH system.9 Depending on the strength of interactions, researchers identified three distinct regimes in the quantum system. In our scenario, we speculate that if many-body interactions play a significant role, the formation of Dirac cones and the critical applied bias values could also be influenced by many-body effects.

In summary, our investigation has provided a comprehensive analysis of field effects on incommensurate quantum systems, focusing on a one-dimensional tight-binding model embedded between two electric contacts and subject to an Aubry–André–Harper potential influenced by an external electric field. Our study amalgamated Anderson and Aubry–André–Harper localization effects on carrier wave functions within our model, further elucidating our findings through the computation of the Lyapunov exponent.

Several intriguing phenomena have been unraveled through our exploration of the Aubry–André–Harper potential. Initially, our observations highlighted a metal–insulator transition occurring at the critical potential strength λc, leading to a profound localization of carrier eigenvectors within the lattice. Notably, we discovered that this critical potential strength, λc, is contingent upon the applied bias value V, diminishing as the external electric field F increases.

In addition, we unveiled the potential to create massless Dirac fermions within the system by carefully adjusting the applied external bias to specific critical values. Consequently, we achieved Dirac crossings within the band structure at defined energies, guiding the dynamics of fermion carriers through a Dirac Hamiltonian characterized by a Fermi velocity, vF. Intriguingly, our exploration showcased that the wave functions of these massless particles can oscillate between extension and localization contingent upon the strength of the Aubry–André–Harper potential, λ. This intriguing property reflects the localization traits inherent in the Aubry–André–Harper eigenstates.

Finally, our study emphasizes the viability of detecting these massless particles through experimental setups. The potential for such an experimental verification opens doors to probing the elusive Aubry–André–Harper metal–insulator transition within tangible materials. Our research, shedding light on carrier behaviors in incommensurate systems, holds significant promise for the broader research domain.

The authors have no conflicts to disclose.

M. Cruz-Méndez: Investigation (lead); Methodology (lead); Software (lead). H. Cruz: Conceptualization (lead); Supervision (lead); Writing – original draft (lead); Writing – review & editing (lead).

Data sharing is not applicable to this article as no new data were created or analyzed in this study.

1.
2.
G.
Roati
,
C.
D’Errico
,
L.
Fallani
,
M.
Fattori
,
C.
Fort
,
M.
Zaccanti
,
G.
Modugno
,
M.
Modugno
, and
M.
Inguscio
,
Nature
453
,
895
(
2008
).
3.
S.
Aubry
and
G.
André
,
Ann. Isr. Phys. Soc.
3
,
133
(
1980
).
4.
P. G.
Harper
,
Proc. Phys. Soc. A
68
,
874
(
1955
).
5.
G. A.
Domínguez-Castro
and
R.
Paredes
,
Eur. J. Phys.
40
,
045403
(
2019
).
6.
D.
Dutta
,
A.
Roy
, and
K.
Saha
,
Phys. Rev. B
107
,
035120
(
2023
).
7.
J.
Billy
,
V.
Josse
,
Z.
Zuo
,
A.
Bernard
,
B.
Hambrecht
,
P.
Lugan
,
D.
Clément
,
L.
Sanchez-Palencia
,
P.
Bouyer
, and
A.
Aspect
,
Nature
453
,
891
(
2008
).
8.
S. S.
Kondov
,
W. R.
McGehee
,
J. J.
Zirbel
, and
B.
DeMarco
,
Science
334
,
66
(
2011
).
9.
K.
Huang
,
D.
Vu
,
X.
Li
, and
S.
Das Sarma
,
Phys. Rev. B
107
,
035129
(
2023
).
10.
J.
Sutradhar
,
S.
Mukerjee
,
R.
Pandit
, and
S.
Banerjee
,
Phys. Rev. B
99
,
224204
(
2019
).
11.
D.
Vu
,
K.
Huang
,
X.
Li
, and
S.
Das Sarma
,
Phys. Rev. Lett.
128
,
146601
(
2022
).
12.
R. B.
Diener
,
G. A.
Georgakis
,
J.
Zhong
,
M.
Raizen
, and
Q.
Niu
,
Phys. Rev. A
64
,
033416
(
2001
).
13.
S.
Xu
,
X.
Li
,
Yi-T.
Hsu
,
B.
Swingle
, and
S.
Das Sarma
,
Phys. Rev. Res.
1
,
032039(R)
(
2019
).
14.
S.
Ray
,
A.
Ghosh
, and
S.
Sinha
,
Phys. Rev. E
97
,
010101(R)
(
2018
).
15.
S.
Ray
,
S.
Sinha
, and
K.
Sengupta
,
Phys. Rev. A
98
,
053631
(
2018
).
16.
M.
Cruz-Méndez
and
H.
Cruz
,
Phys. Lett. A
490
,
129193
(
2023
).
17.
E. J.
Guzmán
,
O.
Oubram
,
O.
Navarro
, and
I.
Rodríguez-Vargas
,
Phys. Rev. B
107
,
045407
(
2023
).
18.
D. N.
Talwar
and
P.
Becla
,
Physica B
650
,
414500
(
2023
).
19.
T.
Yoshihiro
and
N.
Nishiguchi
,
Phys. Rev. B
100
,
235441
(
2019
).
20.
R.
Hu
and
Z.
Tian
,
Phys. Rev. B
103
,
045304
(
2021
).
21.
T.
Juntunen
,
O.
Vänskä
, and
I.
Tittonen
,
Phys. Rev. Lett.
122
,
105901
(
2019
).
23.
M. I.
Katsnelson
,
K. S.
Novoselov
, and
A. K.
Geim
,
Nat. Phys.
2
,
620
(
2006
).
24.
A. F.
Young
and
P.
Kim
,
Nat. Phys.
5
,
222
(
2009
).
25.
N.
Stander
,
B.
Huard
, and
D.
Goldhaber-Gordon
,
Phys. Rev. Lett.
102
,
026807
(
2009
).
26.
V. V.
Cheianov
and
V. I.
Fal’ko
,
Phys. Rev. B
74
,
041403(R)
(
2006
).
27.
J. M.
Pereira
,
V.
Mlinar
,
F. M.
Peeters
, and
P.
Vasilopoulos
,
Phys. Rev. B
74
,
045424
(
2006
).
28.
V. V.
Cheianov
,
V.
Fal’ko
, and
B. L.
Altshuler
,
Science
315
,
1252
(
2007
).
29.
W.-Y.
He
,
Z.-D.
Chu
, and
L.
He
,
Phys. Rev. Lett.
111
,
066803
(
2013
).
30.
G.-H.
Lee
,
G.-H.
Park
, and
H.-J.
Lee
,
Nat. Phys.
11
,
925
(
2015
).
31.
O.
Klein
,
Z. Phys. A: Hadrons Nucl.
53
,
157
(
1929
).
32.
P.
Hewageegana
and
V.
Apalkov
,
Phys. Rev. B
77
,
245426
(
2008
).
33.
A.
Matulis
and
F. M.
Peeters
,
Phys. Rev. B
77
,
115423
(
2008
).
34.
J. S.
Wu
and
M. M.
Fogler
,
Phys. Rev. B
90
,
235402
(
2014
).
35.
J. H.
Bardarson
,
M.
Titov
, and
P. W.
Brouwer
,
Phys. Rev. Lett.
102
,
226803
(
2009
).
36.
C. A.
Downing
,
D. A.
Stone
, and
M. E.
Portnoi
,
Phys. Rev. B
84
,
155437
(
2011
).
37.
Y.
Zhao
,
J.
Wyrick
,
F. D.
Natterer
,
J. F.
Rodriguez-Nieva
,
C.
Lewandowski
,
K.
Watanabe
,
T.
Taniguchi
,
L. S.
Levitov
,
N. B.
Zhitenev
, and
J. A.
Stroscio
,
Science
348
,
672
(
2015
).
38.
J.
Lee
,
D.
Wong
,
J.
Velasco, Jr
,
J. F.
Rodriguez-Nieva
,
S.
Kahn
,
H.-Z.
Tsai
,
T.
Taniguchi
,
K.
Watanabe
,
A.
Zettl
,
F.
Wang
,
L. S.
Levitov
, and
M. F.
Crommie
,
Nat. Phys.
12
,
1032
(
2016
).
39.
C.
Gutierrez
,
L.
Brown
,
C.-J.
Kim
,
J.
Park
, and
A. N.
Pasupathy
,
Nat. Phys.
12
,
1069
(
2016
).
40.
K.-K.
Bai
,
J.-B.
Qiao
,
H.
Jiang
,
H.
Liu
, and
L.
He
,
Phys. Rev. B
95
,
201406(R)
(
2017
).
41.
H.
Cruz
and
S.
Das Sarma
,
J. Phys. I France
3
,
1515
(
1993
).
42.
43.
P.
Cruz
and
H.
Cruz
,
Acta Phys. Pol., A
138
,
21
(
2020
).
44.
D. J.
Boers
,
B.
Goedeke
,
D.
Hinrichs
, and
M.
Holthaus
,
Phys. Rev. A
75
,
063404
(
2007
).
45.
D.
Witthaut
,
T.
Salger
,
S.
Kling
,
C.
Grossert
, and
M.
Weitz
,
Phys. Rev. A
84
,
033601
(
2011
).
46.
H.
Cruz
,
D.
Luis
,
F.
Delgado
, and
J. G.
Muga
,
Phys. Rev. B
70
,
195315
(
2004
).
48.
D.
Luis
,
H.
Cruz
, and
N. E.
Capuj
,
Phys. Rev. B
59
,
9787
(
1999
).
49.
H.
Li
,
Y.-Y.
Wang
,
Y.-H.
Shi
,
K.
Huang
,
X.
Song
,
G.-H.
Liang
,
Z.-Y.
Mei
,
B.
Zhou
,
H.
Zhang
,
J.-C.
Zhang
,
S.
Chen
,
S. P.
Zhao
,
Y.
Tian
,
Z.-Y.
Yang
,
Z.
Xiang
,
K.
Xu
,
D.
Zheng
, and
H.
Fan
,
npj Quantum Inf.
9
,
40
(
2023
).